Electroporation-Based Technologies for Medicine: Principles

BE16CH12-Yarmush
ARI
ANNUAL
REVIEWS
19 June 2014
14:20
Further
Click here for quick links to
Annual Reviews content online,
including:
• Other articles in this volume
• Top cited articles
• Top downloaded articles
• Our comprehensive search
Electroporation-Based
Technologies for Medicine:
Principles, Applications,
and Challenges
Martin L. Yarmush,1,2 Alexander Golberg,1
Gregor Serša,3 Tadej Kotnik,4 and Damijan Miklavčič4
1
Center for Engineering in Medicine, Department of Surgery, Massachusetts General Hospital,
Harvard Medical School and Shriners Burn Hospital for Children, Boston,
Massachusetts 02114; email (M.L.Y.): ireis@sbi.org
2
Department of Biomedical Engineering, Rutgers University, Piscataway, New Jersey 08854;
email: yarmush@rci.rutgers.edu
3
Department of Experimental Oncology, Institute of Oncology Ljubljana, SI-1000 Ljubljana,
Slovenia
4
Department of Biomedical Engineering, Faculty of Electrical Engineering, University of
Ljubljana, SI-1000 Ljubljana, Slovenia; email: damijan.miklavcic@fe.uni-lj.si
Annu. Rev. Biomed. Eng. 2014. 16:295–320
Keywords
First published online as a Review in Advance on
May 27, 2014
pulsed electric field, electropermeabilization, electrochemotherapy, gene
electrotransfer, DNA vaccination, irreversible electroporation
The Annual Review of Biomedical Engineering is
online at bioeng.annualreviews.org
This article’s doi:
10.1146/annurev-bioeng-071813-104622
c 2014 by Annual Reviews.
Copyright All rights reserved
Abstract
When high-amplitude, short-duration pulsed electric fields are applied to
cells and tissues, the permeability of the cell membranes and tissue is increased. This increase in permeability is currently explained by the temporary appearance of aqueous pores within the cell membrane, a phenomenon
termed electroporation. During the past four decades, advances in fundamental and experimental electroporation research have allowed for the translation of electroporation-based technologies to the clinic. In this review, we
describe the theory and current applications of electroporation in medicine
and then discuss current challenges in electroporation research and barriers
to a more extensive spread of these clinical applications.
295
BE16CH12-Yarmush
ARI
19 June 2014
14:20
Contents
1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2. FUNDAMENTAL PRINCIPLES OF ELECTROPORATION . . . . . . . . . . . . . . . . .
2.1. Events at the Molecular Level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2. Transmembrane Voltage and Electroporation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3. Pore Formation, Resealing, and Thermal Damage . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4. Transmembrane Mass Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3. ELECTROPORATION APPLICATIONS IN MEDICINE . . . . . . . . . . . . . . . . . . . . .
3.1. Electrochemotherapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2. Nonthermal Tissue Ablation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3. Gene Therapy and DNA Vaccination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4. Transdermal Drug Delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. TREATMENT PLANNING AND IMAGING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1. Ultrasound . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2. Magnetic Resonance Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3. Computed Tomography and Positron Emission Tomography . . . . . . . . . . . . . . . .
4.4. Electrical Impedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5. POTENTIAL AND CHALLENGES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6. CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Electrochemotherapy
(ECT): a combined
treatment using
electroporation of
membranes and
facilitating transport of
cytotoxic drugs such as
bleomycin and
cisplatin that have a
hindered transport
across the membrane
and intracellular target
(i.e., DNA), thus
potentiating the
cytotoxic effect of such
drugs
Gene electrotransfer
(GET): one
of the gene therapy
modalities using
specific electric pulses
for cellular uptake of
naked plasmid DNA,
encoding for a specific
therapeutic protein
with the means of
molecular targeted
therapy of cancer
296
296
297
297
298
299
301
302
302
303
305
306
307
309
310
311
311
311
312
1. INTRODUCTION
Electroporation is the increase of cell membrane permeability due to externally applied pulsed
electric fields. Although observation of the effects of pulsed electric fields on biological material dates back more than 250 years, only in the past two decades have practical applications of
electroporation emerged in food processing, pharmaceutics, and medicine (1).
The history of electroporation most likely begins in the middle of the eighteenth century
when Nollet (2) reported the first systematic observations of the appearance of red spots on
animal and human skin exposed to electric sparks. Over the next two centuries, Ritter in 1802
(see 3), Frankenhaeuser & Widén (4) in 1956, and Stampfli & Willi (5) in 1957 reported that
nerve membrane electroporation may explain the electrical conductivity changes in nerves that
have been damaged by electric fields. The medical application of electroporation began in 1982
with the seminal work of Neumann and colleagues (6).1 Those authors used pulsed electric fields
to temporarily permeabilize cell membranes to deliver foreign DNA into cells. In the following
decade, the combination of high-voltage pulsed electric fields with the chemotherapeutic drug
bleomycin and with DNA yielded novel clinical applications: electrochemotherapy (ECT) (7–
9) and gene electrotransfer (GET) (10, 11), respectively. In recent years, nonthermal irreversible
electroporation (NTIRE) for the ablation of solid tumors has emerged as a new medical application
of electroporation technology (12, 13).2
1
Neumann and colleagues also coined and first used the term electroporation in 1982.
2
The terms pulsed electric field (PET) treatment, ECT, GET (alternatively defined as gene electrotherapy), electrogene
therapy, electroplasmolysis, and NTIRE are sometimes used interchangeably, but also for better acceptance by consumers.
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
2. FUNDAMENTAL PRINCIPLES OF ELECTROPORATION
2.1. Events at the Molecular Level
Since the discovery of electroporation, several competing theoretical descriptions of the events
underlying the phenomenon have been proposed, assuming either a certain type of deformation of
the lipids (14–16), their phase transition (17), breakdown of interfaces between the domains with
different lipid compositions (18), or denaturation of membrane proteins (19). However, all of these
descriptions suffer from obvious flaws (20), and today there is broad consensus that electroporation
is best described as the formation of aqueous pores in the lipid bilayer (20–23). This description
also explains the prevalent choice of the term electroporation, as opposed to the broader term
electropermeabilization. The latter term describes the consequence rather than the underlying
mechanism and is thus applicable to any of the alternative explanations of the phenomenon as well.
According to the theory of aqueous pore formation, which is based largely on thermodynamic
considerations, formation of aqueous pores is initiated by the penetration of water molecules
into the lipid bilayer of the membrane, which leads to reorientation of the adjacent lipids with
their polar head groups pointing toward these water molecules (Figure 1). Unstable pores with
nanosecond lifetimes can generally form even in the absence of an external electric field, but an
exposure of the membrane to such a field reduces the energy required for penetration of water
into the bilayer. Within the first nanoseconds, this penetration is caused mostly by the transfer
of the external field to the membrane, and then within a microsecond, the transmembrane field
is amplified by polarization, resulting in the buildup of an induced transmembrane voltage (ITV)
(24, 25). The entrance of water increases the probability of pore formation, resulting in a larger
number of pores formed in the bilayer per unit of area and per unit of time, with the pores
a
b
Nonthermal
irreversible
electroporation
(NTIRE): an
electroporation-based
nonthermal tissue
ablation method,
leading to cell death
either directly through
excessive damage
inflicted to cells and
membranes observed
as necrosis or via
apoptosis
c
Figure 1
An idealized molecular-level scheme (top) and an atomic-level molecular dynamics simulation (bottom) of electroporation with the
electric field perpendicular to the bilayer plane. In the simulation, a 1-palmitoyl 2-oleoyl phosphatidylcholine (POPC) bilayer
surrounded by a saline solution is exposed to a field of 4 MV/cm, with the snapshots taken 0, 0.15, and 0.50 ns after the field is turned
on. (a) The intact bilayer. (b) Water molecules start penetrating the bilayer, forming a water wire. (c) The lipids adjacent to the water
wire start reorienting toward the wire with their polar head groups, stabilizing the pore and allowing more water, as well as other polar
molecules and ions, to enter. The atoms of the lipid head groups and tails are shown in orange and gray, respectively; water molecules
in cyan; sodium ions in green; and chloride ions in pink. (Top row reprinted with permission from 181; bottom row reprinted with
permission from 23.)
www.annualreviews.org • Electroporation in Medicine
297
BE16CH12-Yarmush
ARI
19 June 2014
14:20
also being more stable than those formed in the absence of an electric field. For transmembrane
voltages of hundreds of millivolts, the number of pores becomes large enough, and their average
lifetimes long enough (from milliseconds up to minutes), for a detectable increase of membrane
permeability to molecules otherwise unable to cross the membrane.
Aqueous pores in the bilayer have radii of at most several nanometers, which is too small to be
observable by optical microscopy, and the sample preparation required for electron microscopy of
soft matter (vacuumization, fixation, and/or metallic coating) is too harsh for reliable preservation
of semistable structures in the bilayer, such that pores cannot be clearly distinguished from artifacts.
Nevertheless, there is rather convincing evidence in favor of the theory of aqueous pore formation
in the form of molecular dynamics simulations. These simulations largely confirm the hypothesized
view of the sequence of molecular-scale events and also show a clear increase in the rate of pore
formation with an increase in the electric field to which the membrane is exposed—first through
the direct action of the external field, and then augmented by the inducement of transmembrane
voltage resulting from polarization (26).
2.2. Transmembrane Voltage and Electroporation
The exposure of a cell to an external electric field can be achieved either by bringing the cell
into direct contact with the electrodes generating the field (the approach used in the patch clamp
technique) or by placing the cell into the field without such contact. In the latter case, the voltage
formed by the field on the membrane (the ITV) represents only a part of the voltage delivered
to the electrodes; moreover, the ITV varies with position on the membrane, and also with time,
reaching its steady-state value after a gradual buildup.
For regularly shaped cells (close to spheres, spheroids, cylinders, etc.) sufficiently far apart (i.e.,
in dilute suspensions), the spatial distribution and time dependence of the ITV can be derived
analytically and expressed by explicit formulae (27–31). Thus when a single spherical cell with
radius R is exposed to a homogeneous electric field of strength E, the first-order approximation
of the time course of its ITV (sometimes referred to as the Schwan equation) reads
ITV = f ER cos θ(1 − e −t/τ ),
where f is a dimensionless factor, θ is the angle measured from the center of the cell with respect to
the direction of the field, t is the time elapsed since the onset of the field, and τ is the time constant
of membrane charging. [ f and τ can be expressed as functions of the electrical and geometrical
properties of the cell and its surroundings (28, 30), and under physiological conditions, f ≈ 1.5
and τ ≈ 0.5 μs.] The ITV is thus proportional to R and E, and its cosine spatial distribution
implies that the ITV reaches its peak values at the poles of the cell facing the two electrodes and
is equal to zero at the equator of the cell that lies on the cell’s plane of symmetry parallel to the
electrodes.
For irregularly shaped cells, as well as for cells close to each other (in tissues, clusters, and
dense suspensions), the ITV cannot be derived analytically, and numerical methods have to be
used instead (32–34). An alternative to both analytical derivation and numerical computation is
experimental determination of the ITV using a potentiometric dye (35, 36).
As pore formation is governed by statistical thermodynamics (20), it is strictly speaking not a
threshold event, in the sense that the pores would form only at E exceeding a certain fixed value.
Nonetheless, electroporation-mediated transport across the membrane is strongly correlated with
the ITV generated by the external electric field (37), which is, in turn, proportional to this field. The
correlation between the ITV and the electroporation-mediated transport across the membrane can
298
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
be demonstrated particularly clearly by combining potentiometric measurements and monitoring
transmembrane transport on the same cell, with two examples shown in Figure 2.
2.3. Pore Formation, Resealing, and Thermal Damage
There are four general contiguous ranges of the electric field strength (Figure 3) that can be best
characterized with respect to the related molecular transport and the functioning of the cell:
No detectable electroporation: This is the lowest range and the one in which there is no
detectable molecular transport observed.
Reversible electroporation: Here, temporary and limited pathways for molecular transport
are formed, but after the end of the electric pulse, the transport ceases, and the cells remain
viable.
Nonthermal irreversible electroporation: Here, the transport through the membrane—
particularly the leakage of intracellular content—is too extensive and/or the resealing too
slow for the cells to recover, resulting in their death and eventual disintegration; still, there
is generally no thermal damage to the cell nor to the released molecules.
Irreversible electroporation accompanied by thermal effects: In sufficiently strong fields, the
electric currents cause a temperature increase sufficiently high for thermal damage to the
cell as well as to the released molecules.
These ranges depend on the duration of the exposure to the field (i.e., on pulse duration), and
they partly overlap, as pore formation is a stochastic process and as the cells within the population
generally vary in size and orientation with respect to the field direction, which affects the ITV
induced by a given field (for the overlapping of nonthermal and thermal range, see below). The
bounds of these four ranges also vary with the type of the cells exposed. For smaller cells, they
are generally shifted upward—that is, toward higher fields (as smaller cells require higher field
strengths for generation of a given ITV)—and this is also the case for cells enveloped by a wall
(i.e., plants, fungi, algae, bacteria, and some archaea), as poration of this additional layer requires
extra induced voltage. The ranges are also affected by the properties of the medium surrounding
the cells; in particular, the range of thermal effects depends strongly on the electrical conductivity
of the medium, as the density of the electric current generated by a given field is proportional to
this conductivity, and so is the heat dissipated by the current per unit time. Thus the lower bound
of the range of thermal effects is shifted rightward and upward for extracellular conductivities
below the physiological level and in the opposite direction for higher extracellular conductivities,
thus widening or narrowing, respectively, the range of nonthermal irreversible electroporation.
Furthermore, the minimal temperature at which thermal damage occurs is both organism and
molecule dependent. Proteins are generally the most sensitive and start to denature at relatively
small temperature increases (at ∼43–45◦ C in human cells). DNA melting occurs only above
∼70◦ C, whereas most lipids and simpler saccharides are not affected even by boiling.
Similar to pore formation, pore resealing is a stochastic process, but it proceeds on a much
longer timescale. Specifically, the formation of electropores takes nano- to microseconds, whereas
their resealing—as revealed by the return of the membrane’s electric conductivity to its preporation
value and by termination of detectable transmembrane transport—is often completed only within
seconds, or even minutes, after the end of the exposure (38). More detailed measurements reveal
that the resealing proceeds in several stages with time constants ranging from micro- and/or
milliseconds up to tens of seconds (39, 40). Unfortunately, neither the existing theory nor the
experiments can provide a reliable picture of specific events characterizing each of these distinctive
www.annualreviews.org • Electroporation in Medicine
299
ARI
19 June 2014
14:20
a
E
E
0.2
0.3
0.2
0.3
0.1
0.4
0.1
0.4
p=0
p=0
0.5
0.5
0.9
0.6
0.8
0.7
0.9
0.6
5 μm
0.8
5 μm
0.7
E
b
E
0.3
0.5
p=0
0.5
0.9
0.6
0.8
0.1
0.4
0.1
p=0
0.7
0.2
0.3
0.2
0.4
0.9
5 μm
0.6
0.8
5 μm
0.7
c
2,000
650
1,000
0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
p
-325
ITV (mV)
ITV (mV)
325
0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
p
-1,000
-650
-2,000
0
300
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
p
Yarmush et al.
PI fluorescence
(a.u.)
d
PI fluorescence
(a.u.)
BE16CH12-Yarmush
0
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
p
BE16CH12-Yarmush
ARI
19 June 2014
14:20
←−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
Figure 2
The induced transmembrane voltage (ITV) and electroporation of two Chinese hamster ovary (CHO) cells
in a physiological medium: one suspended and almost spherical (left-hand sides of panels a–d ) and the
other attached to a surface and irregularly shaped (right-hand sides of panels a–d ). (a) Changes in the
fluorescence of di-8-ANEPPS, a potentiometric dye reflecting the ITV, with dark and bright regions
corresponding to membrane depolarization and hyperpolarization, respectively. (b) Fluorescence of
propidium iodide (PI), a dye fluorescing only inside the cell and thus reflecting its electroporationmediated influx into the cell. (c) ITV along the path shown in panel a as measured ( green) and as predicted
by numerical computation ( gray). (d ) Fluorescence of PI along the path shown in panel a. The spherical
cell was electroporated by a single 1.5-ms, 650-V/cm pulse, and the attached cell by a single 200-μs,
1,000-V/cm pulse. Additional abbreviations: E, electric field strength; p, normalized arc length along the
membrane. (Figure reprinted with permission from 37.)
stages, and reliable molecular dynamics simulations, even in their most simplified versions (e.g.,
coarse grained), cannot yet cover timescales that extensive.
2.4. Transmembrane Mass Transport
Direct microscopic observations have revealed that the mass transport mechanisms during electroporation depend on the nature of molecules (41, 42). Small charged molecules, for example
propidium iodide (PI), enter the cells from both poles exposed to the electrodes during and after
the application of electric fields (Figure 4). Negatively charged small interfering RNA (siRNA)
molecules, however, transfer into the cell cytoplasm exclusively on the side facing the cathode
only when the electric field is applied (Figure 4). When added after electroporation, siRNA was
unable to silence genes because it did not penetrate the cells, as reported by Paganin-Gioanni et al.
b
105
Fraction of exposed cells (%)
Electric field strength (V cm–1)
a
104
103
102
0
0 10–8
10–7
10–6
10–5
10–4
10–3
100
80
60
40
20
0
10–2
0
250
500
No detectable
electroporation
750
1,000 1,250 1,500 1,750 2,000
Electric field strength (V cm–1)
Exposure duration (s)
Reversible
electroporation
Nonthermal
irreversible
electroporation
Irreversible
electroporation
and thermal effects
Figure 3
Electroporation and thermal effects caused by exposure of cells to electric fields. (a) Reversible electroporation, irreversible
electroporation, and thermal damage as functions of electric field strength and duration. (b) The fractions of nonelectroporated,
reversibly electroporated, and irreversibly electroporated cells as functions of electric field strength, for a fixed exposure duration of
1 ms (i.e., along the blue dashed vertical line in panel a). Note that the field scale is logarithmic in panel a but linear in panel b where it
covers a much narrower range. (Panel a adapted with permission from 182 and 183.)
www.annualreviews.org • Electroporation in Medicine
301
BE16CH12-Yarmush
ARI
19 June 2014
14:20
EP
EGFP
PI
(668 Da)
+
Post EP
Milliseconds–
seconds
PI
Seconds–
minutes
PI
pDNA-POPO3
pDNA-POPO3
siRNA-AF546
siRNA-AF546
10μm
EGFP
pDNA
(~3,100 kDa)
10μm
EGFP
siRNA
(16 kDa)
10μm
–
Figure 4
Direct microscopical observation of electrotransfer of PI, siRNA, and pDNA into murine melanoma cells
(10 pulses, 5 ms, 300 V/cm, 1 Hz). The column on the left shows the EGFP-expressing cell. The
electroporation (center column) in the top row is the first step of the electrotransfer of PI: During
electroporation, penetration can be observed on both sides facing the electrodes. Post electroporation (right
column), diffusion and nuclear labeling can be observed. The electroporation (center column) in the middle
row is the first step of the electrotransfer of pDNA labeled with POPO3: A localized interaction with the
plasma membrane ( fluorescent spots) occurs on the side facing the cathode and continues post electroporation
(right column) before translocation into the cytoplasm (not detected on this timescale). The electroporation
(center column) in the bottom row is the first step of the AF546 siRNA electrotransfer: A rapid and free
penetration into the cytoplasm on the side facing the cathode occurs during the electric pulses. Post
electroporation, the AF546 siRNA is localized diffusely throughout the cytoplasm with no further entry.
Abbreviations: AF546 , an Alexa Fluor dye; EGFP, enhanced green fluorescent protein; EP, electroporation;
pDNA, plasmid DNA; PI, propidium iodide; POPO3, an intercalating dye; siRNA, small interfering RNA.
(Figure adapted with permission from 41.)
(41). By contrast, plasmid DNA (pDNA) penetration begins with the induction of local physical
interactions between the nucleic acid and the cell membrane during the application of electric
fields; it is followed by a slow intracellular release into the cytoplasm after exposure to the electric
field is terminated (42), and then by trafficking to the nucleus. These differences are important
for ECT and GET treatment planning.
3. ELECTROPORATION APPLICATIONS IN MEDICINE
3.1. Electrochemotherapy
Electrochemotherapy (ECT) has reached an established position among local treatments in oncology, both human and veterinary (43, 44). More than 3,000 patients were treated with this
application in the European Union in 2012 (45). The basic principle of ECT is the use of electroporation for delivery of non- or poorly permeant chemotherapeutic drugs into the cells, facilitating
their transmembrane transport with the aim of increasing their cytotoxicity. In preclinical studies, bleomycin, used by Mir et al. (9, 46), and cisplatin, used by Serša et al. (47), were identified
as the most suitable drugs for clinical testing (7, 9, 46–48). The effectiveness of ECT has been
demonstrated on different solid tumor models, with the treatment inducing complete or partial
302
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
responses (49). It should be noted that ECT has an excellent therapeutic index; that is, for antitumor effectiveness, very low drug concentrations of the chemotherapeutics are needed, thus
avoiding toxicity.
During preclinical studies, several mechanisms that contribute to the overall effectiveness of
ECT have been demonstrated. Besides the basic mechanism of increased membrane permeability
leading to increased drug cytotoxicity (50), reduced tumor blood flow (vascular lock) is observed
(51) and results in a vascular disruption effect of ECT (52–55); an enhanced immune response
is also observed (56–58). Thus, combining ECT with an immunostimulatory approach using
recombinant cytokines or GET-delivered plasmids encoding these cytokines (e.g., IL-2, IL-12,
IL-15, GM-CSF, and TNF-α) can give rise to a significant potentiation of ECT treatment (59–62).
The first clinical study on ECT was published in 1991 by Mir et al. (7) and demonstrated the
feasibility, safety, and effectiveness of ECT. This first study stimulated groups in the United States
(Tampa), Slovenia (Ljubljana), and France (Toulouse and Reims) to perform further clinical studies
(63, 64), which were then compiled in a joint paper published in 1998 (65). The coalescence of
the field was marked by the report of a European project called the European Standard Operating
Procedures on Electrochemotherapy (ESOPE) (Figure 5). Based on the results of a clinical study
published in 2006 (66) together with the standard operating procedures (SOP) for ECT using the
electric pulse generator Cliniporator (67), ECT was widely accepted for clinical use throughout
Europe. Clinical indications were published in 2008 (68), and a systematic review and meta-analysis
(69) recently analyzed the results of all the published studies through 2012. Data analysis confirmed
that delivery via ECT has a significantly ( p < 0.001) higher effectiveness (by more than 50%) than
bleomycin or cisplatin injection alone. The overall effectiveness of ECT was 84.1% objective
responses (OR, which include both complete and partial responses), and from these, 59.4% were
complete responses (CR, which indicate complete regressions of tumors after therapy) after a
single ECT treatment. The procedure can, however, be repeated with similar effectiveness (70).
Another recent review and clinical study (71) suggested that the SOP may need refinement, as the
current SOP for ECT may not be suitable for tumors bigger than 3 cm in diameter, which require
multiple, consecutive ECT treatments. Several ongoing ECT studies are targeting superficial
tumors, predominantly melanoma but also chest wall breast cancer recurrences (72–74) and head
and neck cancers (75). Furthermore, the technology is also being adapted for the treatment of
deep-seated tumors, such as colorectal tumors; soft tissue sarcomas; and brain, bone, and liver
metastases (45, 52).
Currently, ECT is being used for treatment of metastases and primary tumors in more than
130 cancer centers in Europe and has been accepted in line with other local tumor treatments
(76). In addition, ECT is also used in veterinary oncology, for treatment of metastases as well as
primary tumors (44, 77–82). The success rate of primary tumor treatment with ECT also provides
good evidence for translation of ECT into treatment of less-advanced tumors in humans.
Vascular lock:
transient reduction of
perfusion, at a greater
extent in tumors than
in normal tissues, that
occurs after
application of electric
pulses to tissues
Vascular disruption
effect of ECT: a
cascade of tumor cell
death, surrounding
obstructed tumor
vessels, resulting from
the drug used in the
electroporation of
endothelial cells and
contributing to the
overall effectiveness of
ECT
3.2. Nonthermal Tissue Ablation
Nonthermal irreversible electroporation (NTIRE) is an emerging minimally invasive surgical
procedure to ablate tissue. In NTIRE, externally applied pulsed electric fields cause irreversible
damage to cells by affecting the cell membrane, while sparing the tissue scaffold, large blood
vessels, and other tissue structures (83–85). The preservation of tissue architecture is an important
and unique property of NTIRE ablation, and it probably contributes to the reduced scarring
observed (as reported in 83, 84, 86).
The first work on NTIRE was published by Davalos et al. (12) in 2005. Using mathematical
models, the authors predicted NTIRE would result in the nonthermal ablation of clinically relevant
www.annualreviews.org • Electroporation in Medicine
303
BE16CH12-Yarmush
ARI
19 June 2014
14:20
Tumor before therapy
Local anesthesia
Application of electric pulses
Injection of the drug
After therapy
Figure 5
Tumor treatment with electrochemotherapy (ECT). Cutaneous or subcutaneous tumor nodules are treated
by ECT either with intratumoral bleomycin or cisplatin administration or with intravenous bleomycin.
Immediately after intratumoral administration or 8 min after intravenous bleomycin administration, electric
pulses are applied using plate or needle electrodes. If the tumor is larger than the gap between the electrodes,
electric pulses are delivered by multiple applications, so that the whole tumor volume is exposed to a
sufficiently high electric field, thereby assuring cell membrane electroporation. (Figure adapted with
permission from 63.)
volumes of tissue. This theoretical work was later confirmed by experimental studies in vitro using
cell cultures (87), small-animal models (88), and large-animal models (89) (Figure 6) and recently
by the first human studies (90, 91). The exact mechanisms of cell death after NTIRE are as
yet unknown, and the reports on pathways that lead to cell death are somewhat contradictory.
For example, Guo et al. (92) reported extensive caspase-3 activation 24 h after NTIRE of rat
hepatocellular carcinoma, which suggests apoptosis. By contrast, José et al. (93) did not detect any
caspase-3-positive cells in the treated area of pancreatic carcinoma. The latter group of authors
argued that the differences observed may be related to the tumor model used; however, whether
NTIRE affects the cell membrane as a final event or induces long-term programmed cell death is
still unresolved. It is important to point out that even at the single-cell level in culture, NTIREinduced cell death is a statistical phenomenon (94, 95). In vivo, only limited attention has been
devoted to the inflammatory response to NTIRE. Al-Sakere et al. (96) concluded that NTIRE does
not induce substantial infiltration of immune cells into the treated tissue, which would suggest
that the immune response is not essential for successful NTIRE. By contrast, Onik et al. (97)
observed an immunologic reaction in the lymph nodes draining the prostate area ablated by
NTIRE. Most recently, Neal et al. (98) showed an improved system response to NTIRE ablation
of renal carcinoma tumors in immunocompetent mice as compared with immunodeficient mice.
Most recent clinical studies have investigated the safety and efficacy of NTIRE treatment of
pancreatic adenocarcinoma (90, 99), renal cell carcinoma (100, 101), lung malignant tumors (101,
102), and hepatic malignant tumors (101, 103, 104). One drawback of NTIRE is that the strong
electric fields induce muscle contractions, therefore requiring special anesthesia (105). Cardiac
arrhythmias developed during NTIRE in several patients have been managed by electrocardiographically synchronized delivery of pulses. Recently, a case report (106) claimed that a patient
304
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
a
d
40x
b
4x
c
40x
e
40x
f
40x
ii
i
Figure 6
Swine liver histopathology after NTIRE. (a) Gross pathologic sectioned specimen of NTIRE-ablated liver
shows areas of discoloration caused by ablation. (b) Preserved large vessel (arrow) in area of NTIRE ablation
stained with H&E. (c) Numerous capillaries (thin arrows) and bile ducts (thick arrows) are stained with vWF,
revealing structural preservation. (d ) The normal liver tissue shows mild vWF staining of vessels and bile
ducts. (e) The ablated zone shows markedly increased vWF-positive staining along the sinusoids, as well as
staining of vessels and bile ducts. ( f ) The ablated zone (i) shows increased apoptotic markers in a TUNEL
assay, compared with normal liver (ii). (Figure adapted with permission from 84.) Abbreviations: H&E,
hematoxylin and eosin; NTIRE, nonthermal irreversible electroporation; TUNEL, terminal
deoxynucleotidyl transferase dUTP nick end labeling; vWF, von Willebrand Factor.
with a pancreatic tumor treated by NTIRE developed melena and hematemesis. The origin of this
complication is still not known, but it may be attributable to proteolytic pancreatic enzyme leakage
resulting from vascular injury during surgery or to still unknown mechanisms related to ablation.
The efficacy of the procedure varies among studies and treated organs. In pancreatic and prostate
cancer, studies have reported 100% success in tumor ablation (90, 91, 99); however, no treated
lung tumors have been successfully ablated to date (101, 102). In liver, the response varied from
50 to 98.1% successful ablation of tumor lesions in different studies (101, 103, 104). In addition to
human studies, NTIRE has also been successfully used for tumor ablation in veterinary medicine
(107–109).
3.3. Gene Therapy and DNA Vaccination
Another biomedical application of electroporation involves gene electrotransfer (GET) to cells
and tissues for gene therapy and DNA vaccination. This technique has been used mainly for DNA
vaccination against infectious diseases, cancer, arthritis, multiple sclerosis, and inflammation following organ transplantation and to genetically modify cells for regenerative medicine applications
(110–113).
In animals, GET was demonstrated first for skin by Titomirov et al. (11), then for liver, muscle,
and tumor tissue (114–116). Since then, GET use has been expanded to other targets, including
brain, kidney, testis, cartilage, arteries, prostate, and cornea (117–119), though transfection efficiency varies considerably across different tissues. Skeletal muscle fibers, which are polynucleated,
are easily accessible and readily transfected, whereas rapidly dividing tumor cells in solid tumors
are the most difficult to transfect (120).
www.annualreviews.org • Electroporation in Medicine
DNA vaccination:
a technique for
protecting an
organism against
disease by injecting it
with genetically
engineered DNA to
produce an
immunological
response more safely
than with viral vectors
305
BE16CH12-Yarmush
ARI
19 June 2014
14:20
In cancer treatment, most of the GET preclinical studies have concentrated on delivering
immunomodulatory genes, such as IL-12, IL-2, IL-15, and GM-CSF (110, 111). In addition,
several studies have documented the use of genes that target the tumor vasculature, such as vasostatin, endostatin, and vascular endothelial growth factor (VEGF) (121–124). Recently, GET of
an antiangiogenic metargidin peptide (AMEP) targeting integrins and GET of siRNA targeting
endoglin have been shown to exert antiproliferative and antiangiogenic effects on melanoma tumors (for AMEP) and mammary carcinoma tumors (for siRNA) (125, 126). Two clinical studies
that have used GET successfully have demonstrated the safety and effectiveness of GET to tumors. The first study, by Daud et al. (10), demonstrated that IL-12 GET to cutaneous melanoma
nodules in humans was very efficient, as tumor necrosis was observed in most of the treated lesions; furthermore, in two of the patients, the response was also observed in distant untreated
metastases. The other clinical study (127), which examined intratumoral GET of plasmid AMEP
in patients with cutaneous melanoma metastases, demonstrated not only the safety of transferring
the molecule into tumors but also some effectiveness.
GET is also being successfully used in veterinary oncology for treatment of tumors in companion animals and in horses. To date, reports on the successful use of IL-12 GET for treatment
of sarcoids in horses and different primary tumors in dogs are available (128, 129). In addition,
one study describes the successful combination of ECT and GET for treatment of tumors in dogs
(130).
GET in DNA vaccination, introduced by Nomura et al. (131) in 1996, aims to address two
major factors that are thought to limit clinical success of human DNA vaccines: (a) low transfection
efficiency and (b) insufficient recruitment of antigen-presenting cells to the injection site (132).
GET has been shown to enhance DNA vaccine efficacy up to three orders of magnitude (132) and
to achieve responses comparable to those achieved with protein vaccines (133). Higher efficiencies
have been obtained with GET than with the gene gun or by direct intramuscular injections (134,
135).
In cell-based regenerative medicine, electroporation is an alternative nonviral method for introducing DNA and RNA into cells. Although viral-mediated transfection is the most efficient
method for genetic manipulation of mammalian cells, virus-related immunogenicity, cytotoxicity, and possible mutations that may disrupt tumor suppressor genes, activate oncogenes, or
interrupt essential genes are major drawbacks of viral methods for clinical applications (136).
Multiple GET protocols have been published for various cell types with commercially available
solutions for optimizing the electroporation procedure (137, 138). Defining the optimum conditions for specific-cell-line electroporation can be a labor- and cost-intensive procedure. To
avoid this problem, a high-throughput method has been introduced to identify electroporation
parameters for cell electroporation in a single step using a concentric electrode configuration
(139).
Microfluidics provides a convenient and cost-effective way to design miniaturized devices to
electroporate large volumes of cells, which has been a challenge for translating cell therapy to
clinical application (140). In contrast to the classic bulk electroporation system, microfluidic platforms can provide a controlled microenvironment for single-cell transfection (141). Recently,
Geng et al. (142) reported a ∼75% transfection efficiency using a microfluidic system with a flow
rate of ∼20 mL/min, which may be sufficient for clinical cell-therapy applications.
3.4. Transdermal Drug Delivery
The barrier property of skin provides obvious survival benefits in the context of variable and
harsh environments (i.e., in the presence of environmental threats of a chemical, physical,
306
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
or biological nature). The skin also undergoes continuous regeneration and repair and is composed of several layers: the stratum corneum, viable epidermis, dermis, and subcutaneous tissue (hypodermis). Owing to its size and accessibility, the skin represents a convenient entry element for drugs. In particular, transdermal delivery has potential for introducing drugs
that are not suitable for intravenous or oral administration, and can overcome many of the
disadvantages of other delivery routes. By using the transdermal route of administration,
we can avoid the liver’s first-pass effect (via hepatic metabolism), obtain better pharmacokinetic profiles, reduce side effects, and achieve reasonable patient compliance. However, simple
transdermal delivery often offers limited flux owing primarily to the impermeable outermost
layer of the skin—the stratum corneum. Overcoming this barrier can be achieved by different enhancement methods, both passive and active (143), one of them being electroporation
(144, 145).
Electroporation has been shown to create aqueous pathways across the stratum corneum and
thus to enhance transdermal drug delivery. The stratum corneum’s electrical resistance is orders
of magnitude higher than that of deeper tissues, and the high electric field resulting from the
application of electric pulses remains mostly in the stratum corneum. As a result of electroporation,
the resistance of the stratum corneum rapidly decreases, and the electric field distributes into
deeper tissue layers (Figure 7) (146, 147).
Exponentially decaying pulses and square wave–shaped pulses are the most commonly used
pulse types in electroporation. In general, increasing the pulse amplitude, duration, and number results in increased permeability of the stratum corneum, allowing for better drug transport
(148). Consequently, molecular flux is enhanced until it reaches a maximum value beyond which
increasing pulse parameters has negligible effect on flux (149). Electroporation has been used
successfully to enhance skin permeability for molecules with differing lipophilicity and size (small
molecules, proteins, peptides, and oligonucleotides), including drugs such as fentanyl, insulin, and
methotrexate (144, 150, 151).
Molecular transport through electroporated skin occurs through different mechanisms, mainly
(a) enhanced diffusion during and after application of electric pulses and (b) electrophoretic movement with very slight electroosmosis during application of a pulse (152). It should be stressed
that reproducibility of drug delivery (i.e., dose control) using electroporation is rather poor and
represents a major obstacle in using the technique as a drug delivery enhancer. Electroporation
can, however, be readily used for applications such as gene delivery. As discussed above, electroporation can be used for applications in which, after intradermal injection, therapeutic molecules
need to be inserted in viable skin cells, a process that is now increasingly being tested for DNA
vaccination (153, 154).
Electroporation can also be used for transdermal delivery of low-molecular-weight peptide
vaccines and, thus, for cancer and infection disease in a type of needleless vaccination (155, 156).
Electric pulses and electrodes must be carefully selected in order to avoid skin damage and/or
irritation as well as pain.
4. TREATMENT PLANNING AND IMAGING
Treatment planning is essential for successful electroporation. Efficient cell membrane electroporation depends on establishing a sufficiently high electric field locally (i.e., in the target tissue).
The ultimate goal of treatment planning is, therefore, to model electrode position and number;
electric field amplitude; and pulse duration, number, and frequency to nonthermally ablate only
the targeted tissue. The current treatment planning models focus on (a) electrode position, to
cover the target tissue by electric fields and spare nontarget tissue (89); (b) electric field protocol
www.annualreviews.org • Electroporation in Medicine
307
BE16CH12-Yarmush
ARI
19 June 2014
a
14:20
Corneocyte
interior
Electric pulses
Corneocyte
interior
Stratum
corneum
b
c
Figure 7
Electroporation of the stratum corneum. (a) Electric pulses create pathways through lipid bilayers of
corneocytes in the stratum corneum, making a transcellular route amenable for drug transport. (b,c) In an
experiment using fresh full-thickness pig ear skin, a patch with 100 μL of 2.5% patent blue solution was
placed on the skin for 5 min before pulse delivery. During skin electroporation treatment, electric pulses of
200 V and 100-μs duration were delivered continuously for 30 s. Peak current during the pulses, measured
using digital oscilloscope, was ∼20 mA. After treatment, the patent blue patch was applied for another
15 min. Tenfold magnification of the histological section of the skin sample is shown (b) after
electroporation treatment and (c) after passive diffusion. (Panel a adapted with permission from 140; panels b
and c reproduced with permission from 184.)
optimization, to delineate the thermal damage from NTIRE and avoid irreversible electroporation
in the case of GET or ECT (157); and (c) integration of mathematically derived treatment planning with diagnostic imaging (158, 159). An example of a treatment plan used in ECT in a case
of liver colorectal metastasis (52) is shown in Figure 8.
Phase and fluorescent microscopic studies have led to a partial understanding of the
mechanisms behind cell membrane electroporation. These studies have shown that different
types of molecules—small molecules, RNA, and DNA—have different transmembrane transport mechanisms during electroporation (41). However, in clinics an important advantage of
electroporation-based technologies is the ability of physicians to use existing imaging methods—
ultrasound (US), magnetic resonance imaging (MRI), computed tomography (CT), and positron
308
Yarmush et al.
ARI
19 June 2014
14:20
a
b
c
Electrode Voltage
pair
(V)
1–5
2–5
3–6
4–6
1–6
2–6
3–5
4–5
5–6
Tumor
Hepatic
veins
Vena cava
1,300
1,700
1,900
1,700
2,100
2,100
2,100
2,100
1,700
Volume fraction of
tumor tissue
BE16CH12-Yarmush
5−6
2−6
1.0
1−6
0.8
4−5
3−5
0.6
3−6
4−6
0.4
2−5
0.2
1−5
0
400
500
600
700
800
Local electric field
strength (V cm–1)
d
e
1
f
2
g
5
5
h
6
6
3
i
1
6
j
2
k
l
5
5
4
m
6
5
6
3
4
Figure 8
Treatment plan used in a case of colorectal liver metastasis (52). The tumor was located between major hepatic veins and vena cava and
was not amenable to surgical resection or RFA, so electroporation was performed using six needle electrodes 20 cm in length with an
insulating sleeve, an active tip of 4 cm, and a diameter of 1.2 mm. The electrodes were introduced using ultrasound during open
surgery, and high-voltage pulses, as given in panel b, were delivered between pairs of electrodes, assuring coverage of the whole tumor
volume with sufficiently high electric field. The required electric field for the tumor was 400 V/cm for eight pulses of 100-μs duration.
Panel c shows the fraction of tumor volume covered by 400+ V/cm produced with each pair of electrodes. Actual placement of the
electrodes during open surgery is shown in panel d. Panels e through m provide graphical representations of the tumor fraction covered
with 400 V/cm in a selected cross section for each pair of electrodes, with the corresponding electrode pairs marked with Arabic
numerals. The area exposed to an electric field higher than 400 V/cm appears dark/brown, whereas the tumor is light/green.
emission tomography (PET)—for procedure guidance. Below, we briefly describe use of these
imaging methods to monitor clinical electroporation effects and mechanisms.
4.1. Ultrasound
US has been used in animal and clinical studies for real-time electrode positioning and observing
both the immediate and posttreatment effects of electroporation. The US findings after electroporation are dynamic (Figure 9a) and tightly correlated with histological observations. In liver,
the irreversibly electroporated region is immediately visualized as a hypoechoic area (Figure 9a,
left) (85); however, after 90–120 min, an external hyperechoic rim appears, probably owing to
hemorrhagic infiltration (160). The hypoechoic area transitioned fully to hyperechoic 24 h after
treatment (85). The tightest correlation between the size of histological findings of ablated tissue
and the hyperechoic rim was observed 90–120 min after NTIRE treatment (160).
www.annualreviews.org • Electroporation in Medicine
309
BE16CH12-Yarmush
ARI
19 June 2014
14:20
a
b
c
d
Figure 9
(a) US imaging of NTIRE of a normal pig liver. The ablated zone (arrows) is shown (left) immediately after
the procedure and (right) 24 h later. (Panel adapted with permission from 85.) (b) MRI of NTIRE of a
human liver. (left) T2-weighted axial MRI of the liver 2 days post NTIRE shows a hyperintensive reactive
rim (bold arrow) surrounding the ablated region with a hypointense necrotic center (thin arrow). (center)
T1-weighted axial MRI 2 days post NTIRE demonstrates mixed signal intensity (arrow) composed of a
hypointense ablated zone containing a hyperintense central signal. (right) Post-contrast-enhanced
hepatic-arterial-phase axial T1-weighted MRI shows a nonenhancing linear ablation zone and
hyperenhancing peripheral rim (arrow), which corresponds with the hyperemic rim in the left-hand side of
panel b. (Panel adapted with permission from 165.) (c) CT and PET imaging of NTIRE of a human liver.
(left) A contrast-enhanced CT image shows the ablation zone (arrow) immediately after NTIRE. (center) The
peripheral zone of the NTIRE-ablated region (arrow) shows increased FDG uptake 3 days post NTIRE.
(right) At 1 month post NTIRE, a PET image does not show increased FDG uptake within the ablated zone
or at the hyperemic reactive rim. (Panel adapted with permission from 165.) (d ) CT and PET imaging of
ECT of face melanoma. ( far left) A pretreatment CT scan with evidence of a sandglass-shaped mass (arrow)
on the right cheek. (center left) An FDG-PET scan shows an SUVmax of 19.5. (center right) Post-ECT
FDG-PET scan shows that the SUVmax decreased to 5. ( far right) Post-ECT FDG-PET scan showing the
SUVmax having decreased to 1.3. (Panel adapted with permission from 166.) Abbreviations: CT, computed
tomography; ECT, electrochemotherapy; FDG, fluorodeoxyglucose; MRI, magnetic resonance imaging;
NTIRE, nonthermal irreversible electroporation; PET, positron emission tomography; SUVmax, maximum
standardized uptake value; US, ultrasound.
4.2. Magnetic Resonance Imaging
MRI techniques are often used for detecting changes in tissue structure and assessing functionality
after ECT and NTIRE in animals and in human subjects. MRI imaging of NTIRE is shown in
Figure 9b. Systematic MRI studies of NTIRE-ablated tissue demonstrated dynamic behavior
of various MRI sequences detected in the ablated tissue. Indeed, studies have shown that the
ablated volumes detected by T1-weighted (T1W), transverse relaxation time (T2), T2-weighted
(T2W), fluid attenuated inversion recovery (FLAIR), and diffusion-weighted magnetic resonance
imaging (DW-MRI), for the apparent water diffusion coefficient (ADC), are similar to the volumes
detected by histological observations. Additionally, T1-Thrive5-3D-GE urogram scans, together
with intravenous urography, were used to validate the functionality of NTIRE-ablated swine
kidney (161). The usefulness of a particular MRI sequence depends on the type of ablated tissue.
In a murine model, although the ADC allowed the visualization of early and rapid changes in
ECT-treated tumors (162), no significant changes in ADC were observed during blood-brainbarrier NTIRE ablation up to 30 min post treatment (163). Contrast agents such as Gd-DOTA
have already been successfully used for the purpose of observing reversibly electroporated areas
either in muscle tissues or in the brain (163, 164).
310
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
4.3. Computed Tomography and Positron Emission Tomography
Non-contrast- and contrast-enhanced CT have been used in animal experiments and in patient
care for diagnosis, electrode positioning, and posttreatment evaluation of tissue ablation and tumor
regression. NTIRE ablation of liver results in a hypoattenuating area with a hyperattenuating rim
in the ablated area 2 days after the procedure (85) (Figure 9b, left). Contrast-enhanced CT was
used to evaluate the treatment response of liver metastases treated by intraoperative ECT. The
response was deemed complete when the treatment zone appeared as a well-defined area of low
attenuation without enhancement (52).
PET scans of NTIRE ablations show a time-dependent response (Figure 9c). At 3 days post
NTIRE, a fludeoxyglucose (FDG) enhancement in the peripheral zone surrounding the ablated
region is present (Figure 9c, center), but the peripheral increase in FDG uptake disappears within
1 month of treatment (Figure 9c, right). (165). Other work with PET showed the effects of ECT
on cheek melanoma (Figure 9d ) (166). PET CT was also used to evaluate the responses of large
tumors (e.g., recurrent breast cancer) being treated with ECT (74).
4.4. Electrical Impedance
In tissues, electroporation leads to an immediately detectable tissue impedance decrease, which can
be used for outcome assessment (167–170). Control of in situ muscle electroporation to reduce cell
damage to cells while assuring successful gene transfection was demonstrated using current and
voltage measurement during pulse delivery (167). Changes in tissue conductivity can be measured
in real time using a galvanic electroporation cell (168, 171). In this approach, electroporation is
delivered via electrodes with different electrochemical potential (e.g., Zn and Cu). The galvanic
current generated by the electrodes and the treated tissue provides direct measurement of electric
impedance changes during and after electroporation without using additional devices (168, 171).
The two-dimensional reconstruction of tissue impedance, electrical impedance tomography (EIT),
in principle allows near real-time monitoring of tissue electroporation (169, 172). In addition to
EIT, magnetic resonance electrical impedance tomography (MREIT) and current density imaging
(CDI) have been proposed for measurement of current and electric field distributions in tissue
during the treatment (173, 174), which would allow immediate corrective action in performing
electroporation-based treatments, thus further improving their efficacy.
5. POTENTIAL AND CHALLENGES
Electroporation increases cell membrane permeability through a nonthermal, chemical-free path.
Therefore, it can be used to (a) reprogram cell and tissue function by introducing external
molecules that affect different cellular pathways, (b) load cells with new materials, and (c) cause cell
damage and death. Accessible imaging of the treatment and its effects, together with fast patient recovery, has generated great interest in the basic and translational aspects of this technique, spurring
the emergence of data across species, from various tissues, and from various clinical applications.
As described above, encouraging results have been reported in animal models and in several clinical trials using ECT, GET, and NTIRE. However, the inability to ablate lung tumors and to
induce long-term immune protection with DNA vaccines, as well as the somewhat contradictory
results with cell therapy, must be clarified and overcome. These problems could potentially be
understood through better integrating imaging for treatment planning and by using optimization
algorithms for electrode positioning. Development and integration of standard electroporation
treatment planning methods are essential for therapy success (175). This integration will provide
more precise treatment with fewer side effects. In addition, new treatment planning methods will
www.annualreviews.org • Electroporation in Medicine
311
BE16CH12-Yarmush
ARI
19 June 2014
14:20
allow for precise dose control, which will lead to the destruction of large, deep-seated tumors by
ECT and NTIRE. Future systematic studies are required to characterize electroporation-induced
dynamic changes in various tissue types.
To use electroporation efficiently, we must understand the basic mechanisms of damage and
regeneration at the molecular, cellular, and tissue levels. Presumably, molecular understanding will
emerge from advanced molecular dynamics simulations and the like. Investigation of the impact
of electroporation on homogeneous and, more importantly, heterogeneous tissues will address
the critical question of selectivity: Can electroporation selectively target specific cell types? In
addition, further studies on the role of the immune system response to electroporation-induced
permeabilization are needed and will presumably provide answers concerning in vivo mechanisms
of therapeutic efficacy. The combined use of electroporation and immune response stimulation
may yield an even more attractive treatment approach.
An additional challenge involves the elimination of electroporation-induced pain and muscle
contraction without using total-body-paralysis agents. Solving this problem will significantly simplify the use of electroporation in medicine, including its use in developing countries. Recently,
a high-frequency pulse delivery regime with a special electrode configuration was proposed for
NTIRE tissue ablation with reduced muscle contraction (176, 177).
Tissue decellularization using irreversible electroporation and perfusion is a promising technology for both in vitro and in vivo regenerative medicine. Although the available data show that
electroporation speeds the decellularization process in vitro (178), a complete analysis of the biological scaffold that is produced is still missing. The host response to the electroporation-prepared
matrix is also still not known.
The clinical data published thus far on electroporation-based applications have been quite
encouraging. Reports describing the clinical advantages of ECT, NTIRE, and electroporationbased DNA vaccination continue to stimulate current basic and applied research in the field.
GET and other nonviral gene delivery systems may gain more attention, as viral vectors have not
proved to be optimal, owing to safety issues such as insertional mutagenesis and immunological
interference (119). The transfection efficiency of GET in different tissues is still low; therefore,
some attempts have been made to improve it either by scavenging reactive oxygen species (ROS)
produced by electroporation (179) or by degrading the extracellular matrix (180).
Although studies have reported electroporation to be safe, additional large-scale-outcomes
research is needed for various clinical conditions, as are comparative studies with other therapeutic approaches. Long-term observations on tumor recurrence, tissue repair, and long-term
immunization will contribute to greater acceptance of electroporation in medical practice.
6. CONCLUSION
Electroporation is a multidisciplinary platform technology with multiple medical applications.
Successful application of electroporation technologies requires close collaboration between physicians, life and computer scientists, and engineers. Current knowledge of electroporation suggests
that it has vast potential for a number of important global medical challenges, including cancer
treatment, infection disease treatment, and vaccination. Although a full understanding of the fundamental mechanisms at the cellular and tissue levels has yet to be developed, we will undoubtedly
see an increase in medical applications of electroporation over the coming years.
DISCLOSURE STATEMENT
D.M. holds patents on electrochemotherapy that have been licensed to IGEA S.p.A. D.M. also
consults for IGEA.
312
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
ACKNOWLEDGMENTS
Part of this research was supported under various grants from the Slovenian Research Agency and
EU Framework Programme and was conducted within the scope of LEA EBAM. This work was
also supported in part by a ECOR Postdoctoral Fellowship Award to A.G. and Shriners Foundation
Grant #85120-BOS. This manuscript is a result of a networking effort of COST TD1104 Action
(www.electroporation.net). D.M. thanks Dr. Bor Kos from University of Ljubljana, Faculty of
Electrical Engineering, for preparing Figure 8.
LITERATURE CITED
1. Miklavčič D. 2012. Network for development of electroporation-based technologies and treatments:
COST TD1104. J. Membr. Biol. 245:591–98
2. Nollet JA. 1754. Recherches sur les causes particulieres des phénoménes électriques. Paris: Guerin & Delatour
3. Noad HM. 1849. Lectures on Electricity: Comprising Galvinism, Magnetism, Electromagnetism, Magneto- and
Thermo-Electricity, and Electro-Physiology. London: Knight. 3rd ed.
4. Frankenhaeuser B, Widén L. 1956. Anode break excitation in desheathed frog nerve. J. Physiol. 131:243–
47
5. Stampfli R, Willi M. 1957. Membrane potential of a Ranvier node measured after electrical destruction
of its membrane. Experientia 13:297–98
6. Neumann E, Schaefer-Ridder M, Wang Y, Hofschneider PH. 1982. Gene transfer into mouse lyoma
cells by electroporation in high electric fields. EMBO J. 1:841–45
7. Mir LM, Belehradek M, Domenge C, Orlowski S, Poddevin B, et al. 1991. [Electrochemotherapy, a new
antitumor treatment: first clinical trial]. C. R. Acad. Sci. III 313:613–18
8. Okino M, Mohri H. 1987. Effects of a high-voltage electrical impulse and an anticancer drug on in vivo
growing tumors. Jpn. J. Cancer Res. Gann 78:1319–21
9. Orlowski S, Belehradek J Jr, Paoletti C, Mir LM. 1988. Transient electropermeabilization of cells in
culture: increase of the cytotoxicity of anticancer drugs. Biochem. Pharmacol. 37:4727–33
10. Daud AI, DeConti RC, Andrews S, Urbas P, Riker AI, et al. 2008. Phase I trial of interleukin-12 plasmid
electroporation in patients with metastatic melanoma. J. Clin. Oncol. 26:5896–903
11. Titomirov AV, Sukharev S, Kistanova E. 1991. In vivo electroporation and stable transformation of skin
cells of newborn mice by plasmid DNA. Biochim. Biophys. Acta 1088:131–34
12. Davalos R, Mir LM, Rubinsky B. 2005. Tissue ablation with irreversible electroporation. Ann. Biomed.
Eng. 33:223–31
13. Golberg A, Yarmush ML. 2013. Nonthermal irreversible electroporation: fundamentals, applications,
and challenges. IEEE Trans. Biomed. Eng. 60:707–14
14. Michael DH, O’Neill ME. 1970. Electrohydrodynamic instability in plane layers of fluid. J. Fluid Mech.
41:571–80
15. Crowley JM. 1973. Electrical breakdown of bimolecular lipid membranes as an electromechanical instability. Biophys. J. 13:711–24
16. Steinchen A, Gallez D, Sanfeld A. 1982. A viscoelastic approach to the hydrodynamic stability of membranes. J. Colloid Interface Sci. 85:5–15
17. Sugár IP. 1979. A theory of the electric field-induced phase transition of phospholipid bilayers. Biochim.
Biophys. Acta 556:72–85
18. Cruzeiro-Hansson L, Mouritsen OG. 1988. Passive ion permeability of lipid membranes modelled via
lipid-domain interfacial area. Biochim. Biophys. Acta 944:63–72
19. Tsong TY. 1991. Electroporation of cell membranes. Biophys. J. 60:297–306
20. Weaver JC, Chizmadzhev YA. 1996. Theory of electroporation: a review. Bioelectrochem. Bioenerg.
41:135–60
21. Spugnini EP, Arancia G, Porrello A, Colone M, Formisano G, et al. 2007. Ultrastructural modifications
of cell membranes induced by electroporation on melanoma xenografts. Microsc. Res. Tech. 70:1041–50
22. Freeman SA, Wang MA, Weaver JC. 1994. Theory of electroporation of planar bilayer membranes:
predictions of the aqueous area, change in capacitance, and pore-pore separation. Biophys. J. 67:42–56
www.annualreviews.org • Electroporation in Medicine
313
BE16CH12-Yarmush
ARI
19 June 2014
14:20
23. Kotnik T, Kramar P, Pucihar G, Miklavčič D, Tarek M. 2012. Cell membrane electroporation—part 1:
the phenomenon. IEEE Electrical Insulation Mag. 28:14–23
24. Schoenbach K, Beebe S, Buescher E. 2001. Intracellular effect of ultrashort electrical pulses. Bioelectromagnetics 22:440–48
25. Kotnik T, Miklavčič D. 2006. Theoretical evaluation of voltage inducement on internal membranes of
biological cells exposed to electric fields. Biophys. J. 90:480–91
26. Delemotte L, Tarek M. 2012. Molecular dynamics simulations of lipid membrane electroporation.
J. Membr. Biol. 245:531–43
27. Gimsa J, Wachner D. 2001. Analytical description of the transmembrane voltage induced on arbitrarily
oriented ellipsoidal and cylindrical cells. Biophys. J. 81:1888–96
28. Kotnik T, Bobanović F, Miklavčič D. 1997. Sensitivity of transmembrane voltage induced by applied
electric fields—a theoretical analysis. Bioelectrochem. Bioenerg. 43:285–91
29. Kotnik T, Miklavčič D. 2000. Analytical description of transmembrane voltage induced by electric fields
on spheroidal cells. Biophys. J. 79:670–79
30. Kotnik T, Miklavčič D. 2000. Second-order model of membrane electric field induced by alternating
external electric fields. IEEE Trans. Biomed. Eng. 47:1074–81
31. Kotnik T, Miklavčič D, Slivnik T. 1998. Time course of transmembrane voltage induced by time-varying
electric fields—a method for theoretical analysis and its application. Bioelectrochem. Bioenerg. 45:3–16
32. Pucihar G, Kotnik T, Valič B, Miklavčič D. 2006. Numerical determination of transmembrane voltage
induced on irregularly shaped cells. Ann. Biomed. Eng. 34:642–52
33. Ying WJ, Henriquez CS. 2007. Hybrid finite element method for describing the electrical response of
biological cells to applied fields. IEEE Trans. Biomed. Eng. 54:611–20
34. Pucihar G, Miklavčič D, Kotnik T. 2009. A time-dependent numerical model of transmembrane voltage
inducement and electroporation of irregularly shaped cells. IEEE Trans. Biomed. Eng. 56:1491–501
35. Loew LM. 1992. Voltage sensitive dyes: measurement of membrane potentials induced by DC and AC
electric fields. Bioelectromagnetics 13:179–89
36. Pucihar G, Kotnik T, Miklavčič D. 2009. Measuring the induced membrane voltage with di-8-ANEPPS.
J. Visualized Exp. 33:1659
37. Kotnik T, Pucihar G, Miklavčič D. 2010. Induced transmembrane voltage and its correlation with
electroporation-mediated molecular transport. J. Membr. Biol. 236:3–13
38. Saulis G, Venslauskas MS, Naktinis J. 1991. Kinetics of pore resealing in cell membranes after electroporation. Bioelectrochem. Bioenerg. 26:1–13
39. Hibino M, Itoh H, Kinosita K. 1993. Time courses of cell electroporation as revealed by submicrosecond
imaging of transmembrane potential. Biophys. J. 64:1789–800
40. Pucihar G, Kotnik T, Miklavčič D, Teissié J. 2008. Kinetics of transmembrane transport of small
molecules into electropermeabilized cells. Biophys. J. 95:2837–48
41. Paganin-Gioanni A, Bellard E, Escoffre JM, Rols MP, Teissié J, Golzio M. 2011. Direct visualization at
the single-cell level of siRNA electrotransfer into cancer cells. Proc. Natl. Acad. Sci. USA 108:10443–47
42. Golzio M, Teissié J, Rols M-P. 2002. Direct visualization at the single-cell level of electrically mediated
gene delivery. Proc. Natl. Acad. Sci. USA 99:1292–97
43. Testori A, Tosti G, Martinoli C, Spadola G, Cataldo F, et al. 2010. Electrochemotherapy for cutaneous
and subcutaneous tumor lesions: a novel therapeutic approach. Dermatol. Ther. 23:651–61
44. Čemažar M, Tamzali Y, Serša G, Tozon N, Mir LM, et al. 2008. Electrochemotherapy in veterinary
oncology. J. Vet. Intern. Med. 22:826–31
45. Miklavčič D, Serša G, Brecelj E, Gehl J, Soden D, et al. 2012. Electrochemotherapy: technological
advancements for efficient electroporation-based treatment of internal tumors. Med. Biol. Eng. Comput.
50:1213–25
46. Mir LM, Orlowski S, Belehradek J Jr, Paoletti C. 1991. Electrochemotherapy potentiation of antitumour
effect of bleomycin by local electric pulses. Eur. J. Cancer 27:68–72
47. Serša G, Čemažar M, Miklavčič D. 1995. Antitumor effectiveness of electrochemotherapy with cisdiamminedichloroplatinum(II) in mice. Cancer Res. 55:3450–55
48. Gehl J, Skovsgaard T, Mir LM. 1998. Enhancement of cytotoxicity by electropermeabilization: an
improved method for screening drugs. Anticancer Drugs 9:319–25
314
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
49. Gehl J. 2003. Electroporation: theory and methods, perspectives for drug delivery, gene therapy and
research. Acta Physiol. Scand. 177:437–47
50. Čemažar M, Miklavčič D, Ščančar J, Dolžan V, Golouh R, Serša G. 1999. Increased platinum accumulation in SA-1 tumour cells after in vivo electrochemotherapy with cisplatin. Br. J. Cancer 79:1386–91
51. Bellard E, Markelc B, Pelofy S, Le Guerroue F, Serša G, et al. 2012. Intravital microscopy at the single
vessel level brings new insights of vascular modification mechanisms induced by electropermeabilization.
J. Control. Release 163:396–403
52. Edhemović I, Gadžijev EM, Brecelj E, Miklavčič D, Kos B, et al. 2011. Electrochemotherapy: a new
technological approach in treatment of metastases in the liver. Technol. Cancer Res. Treat. 10:475–85
53. Jarm T, Čemažar M, Miklavčič D, Serša G. 2010. Antivascular effects of electrochemotherapy: implications in treatment of bleeding metastases. Expert Rev. Anticancer Ther. 10:729–46
54. Markelc B, Serša G, Čemažar M. 2013. Differential mechanisms associated with vascular disrupting
action of electrochemotherapy: intravital microscopy on the level of single normal and tumor blood
vessels. PLoS ONE 8:e59557
55. Serša G, Jarm T, Kotnik T, Coer A, Podkrajšek M, et al. 2008. Vascular disrupting action of electroporation and electrochemotherapy with bleomycin in murine sarcoma. Br. J. Cancer 98:388–98
56. Roux S, Bernat C, Al-Sakere B, Ghiringhelli F, Opolon P, et al. 2008. Tumor destruction using electrochemotherapy followed by CpG oligodeoxynucleotide injection induces distant tumor responses.
Cancer Immunol. Immunother. 57:1291–300
57. Sedlar A, Dolinšek T, Markelc B, Prosen L, Kranjc S, et al. 2012. Potentiation of electrochemotherapy
by intramuscular IL-12 gene electrotransfer in murine sarcoma and carcinoma with different immunogenicity. Radiol. Oncol. 46:302–11
58. Serša G, Miklavčič D, Čemažar M, Belehradek J, Jarm T, Mir LM. 1997. Electrochemotherapy with
CDDP on LPB sarcoma: comparison of the anti-tumor effectiveness in immunocompetent and immunodeficient mice. Bioelectrochem. Bioenerg. 43:279–83
59. Heller L, Pottinger C, Jaroszeski MJ, Gilbert R, Heller R. 2000. In vivo electroporation of plasmids
encoding GM-CSF or interleukin-2 into existing B16 melanomas combined with electrochemotherapy
induces long-term antitumour immunity. Melanoma Res. 10:577–83
60. Serša G, Čemažar M, Menart V, Gaberc-Porekar V, Miklavčič D. 1997. Anti-tumor effectiveness of
electrochemotherapy with bleomycin is increased by TNF-α on SA-1 tumors in mice. Cancer Lett.
116:85–92
61. Torrero MN, Henk WG, Li S. 2006. Regression of high-grade malignancy in mice by bleomycin and
interleukin-12 electrochemogenetherapy. Clin. Cancer Res. 12:257–63
62. Mir LM, Orlowski S, Poddevin B, Belehradek J Jr. 1992. Electrochemotherapy tumor treatment is
improved by interleukin-2 stimulation of the host’s defenses. Eur. Cytokine Netw. 3:331–34
63. Heller R. 1995. Treatment of cutaneous nodules using electrochemotherapy. J. Fla. Med. Assoc. 82:147–
50
64. Serša G, Štabuc B, Čemažar M, Miklavčič D, Rudolf Z. 2000. Electrochemotherapy with cisplatin:
clinical experience in malignant melanoma patients. Clin. Cancer Res. 6:863–67
65. Mir LM, Glass LF, Serša G, Teissié J, Domenge C, et al. 1998. Effective treatment of cutaneous and
subcutaneous malignant tumours by electrochemotherapy. Br. J. Cancer 77:2336–42
66. Marty M, Serša G, Garbay JR, Gehl J, Collins CG, et al. 2006. Electrochemotherapy—an easy, highly
effective and safe treatment of cutaneous and subcutaneous metastases: results of ESOPE (European
Standard Operating Procedures of Electrochemotherapy) study. Eur. J. Cancer Suppl. 4:3–13
67. Mir LM, Gehl J, Serša G, Collins CG, Garbay J-R, et al. 2006. Standard operating procedures of the
electrochemotherapy: Instructions for the use of bleomycin or cisplatin administered either systemically
or locally and electric pulses delivered by the CliniporatorTM by means of invasive or non-invasive
electrodes. Eur. J. Cancer Suppl. 4:14–25
68. Serša G, Miklavčič D, Čemažar M, Rudolf Z, Pucihar G, Snoj M. 2008. Electrochemotherapy in treatment of tumours. Eur. J. Surg. Oncol. 34:232–40
69. Mali B, Jarm T, Snoj M, Serša G, Miklavčič D. 2013. Antitumor effectiveness of electrochemotherapy:
a systematic review and meta-analysis. Eur. J. Surg. Oncol. 39:4–16
www.annualreviews.org • Electroporation in Medicine
315
BE16CH12-Yarmush
ARI
19 June 2014
14:20
70. Campana LG, Mocellin S, Basso M, Puccetti O, De Salvo GL, et al. 2009. Bleomycin-based electrochemotherapy: clinical outcome from a single institution’s experience with 52 patients. Ann. Surg.
Oncol. 16:191–99
71. Mali B, Miklavčič D, Campana LG, Čemažar M, Serša G, et al. 2013. Tumor size and effectiveness of
electrochemotherapy. Radiol. Oncol. 47:32–41
72. Serša G, Čufer T, Paulin SM, Čemažar M, Snoj M. 2012. Electrochemotherapy of chest wall breast
cancer recurrence. Cancer Treat. Rev. 38:379–86
73. Campana LG, Valpione S, Falci C, Mocellin S, Basso M, et al. 2012. The activity and safety of electrochemotherapy in persistent chest wall recurrence from breast cancer after mastectomy: a Phase-II
study. Breast Cancer Res. Treat. 134:1169–78
74. Matthiessen LW, Johannesen HH, Hendel HW, Moss T, Kamby C, Gehl J. 2012. Electrochemotherapy
for large cutaneous recurrence of breast cancer: a Phase II clinical trial. Acta Oncol. 51:713–21
75. Gargiulo M, Papa A, Capasso P, Moio M, Cubicciotti E, Parascandolo S. 2012. Electrochemotherapy
for non-melanoma head and neck cancers: clinical outcomes in 25 patients. Ann. Surg. 255:1158–64
76. Testori A, Faries MB, Thompson JF, Pennacchioli E, Deroose JP, et al. 2011. Local and intralesional
therapy of in-transit melanoma metastases. J. Surg. Oncol. 104:391–96
77. Kodre V, Čemažar M, Pečar J, Serša G, Cor A, Tozon N. 2009. Electrochemotherapy compared to
surgery for treatment of canine mast cell tumours. In Vivo 23:55–62
78. Mir LM, Devauchelle P, Quintin-Colonna F, Delisle F, Doliger S, et al. 1997. First clinical trial of cat
soft-tissue sarcomas treatment by electrochemotherapy. Br. J. Cancer 76:1617–22
79. Rols MP, Tamzali Y, Teissié J. 2002. Electrochemotherapy of horses: a preliminary clinical report.
Bioelectrochemistry 55:101–5
80. Spugnini EP, Vincenzi B, Citro G, Dotsinsky I, Mudrov T, Baldi A. 2011. Evaluation of cisplatin as an
electrochemotherapy agent for the treatment of incompletely excised mast cell tumors in dogs. J. Vet.
Intern. Med. 25:407–11
81. Tamzali Y, Borde L, Rols MP, Golzio M, Lyazrhi F, Teissié J. 2012. Successful treatment of equine
sarcoids with cisplatin electrochemotherapy: a retrospective study of 48 cases. Equine Vet. J. 44:214–20
82. Tozon N, Serša G, Čemažar M. 2001. Electrochemotherapy: potentiation of local antitumour effectiveness of cisplatin in dogs and cats. Anticancer Res. 21:2483–88
83. Maor E, Ivorra A, Leor J, Rubinsky B. 2007. The effect of irreversible electroporation on blood vessels.
Technol. Cancer Res. Treat. 6:307–12
84. Phillips MA, Narayan R, Padath T, Rubinsky B. 2012. Irreversible electroporation on the small intestine.
Br. J. Cancer 106:490–95
85. Lee EW, Chen C, Prieto VE, Dry SM, Loh CT, Kee ST. 2010. Advanced hepatic ablation technique
for creating complete cell death: irreversible electroporation. Radiology 255:426–33
86. Golberg A, Broelsch G, Bohr S, Mihm M, Austen W, et al. 2013. Non-thermal, pulsed electric field cell
ablation: a novel tool for regenerative medicine and scarless skin regeneration. Technology 1:1–8
87. Miller L, Leor J, Rubinsky B. 2005. Cancer cells ablation with irreversible electroporation. Technol.
Cancer Res. Treat. 4:699–705
88. Edd JF, Horowitz L, Davalos RV, Mir LM, Rubinsky B. 2006. In vivo results of a new focal tissue
ablation technique: irreversible electroporation. IEEE Trans. Biomed. Eng. 53:1409–15
89. Rubinsky B, Onik G, Mikus P. 2007. Irreversible electroporation: a new ablation modality—clinical
implications. Technol. Cancer Res. Treat. 6:37–48
90. Martin RCG, McFarland K, Ellis S, Velanovich V. 2012. Irreversible electroporation therapy in the
management of locally advanced pancreatic adenocarcinoma. J. Am. Coll. Surg. 215:361–69
91. Onik G, Rubinsky B. 2010. Irreversible electroporation: first patient experience focal therapy of prostate
cancer. In Irreversible Electroporation, ed. B Rubinsky, pp. 235–47. Berlin: Springer
92. Guo Y, Zhang Y, Klein R, Nijm GM, Sahakian AV, et al. 2010. Irreversible electroporation therapy in
the liver: longitudinal efficacy studies in a rat model of hepatocellular carcinoma. Cancer Res. 70:1555–63
93. José A, Sobrevals L, Ivorra A, Fillat C. 2012. Irreversible electroporation shows efficacy against pancreatic
carcinoma without systemic toxicity in mouse models. Cancer Lett. 317:16–23
94. Golberg A, Bei M, Sheridan R, Yarmush M. 2013. Regeneration and control of human fibroblast cell
density by intermittently delivered pulsed electric fields. Biotechnol. Bioeng. 110:1759–68
316
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
95. Golberg A, Rubinsky B. 2010. A statistical model for multidimensional irreversible electroporation cell
death in tissue. Biomed. Eng. OnLine 9:13
96. Al-Sakere B, Bernat C, André F, Connault E, Opolon P, et al. 2007. A study of the immunological
response to tumor ablation with irreversible electroporation. Technol. Cancer Res. Treat. 6:301–5
97. Onik G, Mikus P, Rubinsky B. 2007. Irreversible electroporation: implications for prostate ablation.
Technol. Cancer Res. Treat. 6:295–300
98. Neal RE II, Rossmeisl JH Jr, Robertson JL, Arena CB, Davis EM, et al. 2013. Improved local and systemic
anti-tumor efficacy for irreversible electroporation in immunocompetent versus immunodeficient mice.
PLoS ONE 8:e64559
99. Bagla S, Papadouris D. 2012. Percutaneous irreversible electroporation of surgically unresectable pancreatic cancer: a case report. J. Vasc. Interv. Radiol. 23:142–45
100. Pech M, Janitzky A, Wendler JJ, Strang C, Blaschke S, et al. 2011. Irreversible electroporation of renal
cell carcinoma: a first-in-man Phase I clinical study. Cardiovasc. Interv. Radiol. 34:132–38
101. Thomson KR, Cheung W, Ellis SJ, Federman D, Kavnoudias H, et al. 2011. Investigation of the safety
of irreversible electroporation in humans. J. Vasc. Interv. Radiol. 22:611–21
102. Usman M, Moore W, Talati R, Watkins K, Bilfinger TV. 2012. Irreversible electroporation of lung
neoplasm: a case series. Med. Sci. Monit. 18:CS43–47
103. Bruix J, Izzo F, Crocetti L, Vilgrain V, Abdel-Rehim M, et al. 2012. Irreversible electroporation for
the treatment of early-stage hepatocellular carcinoma: a prospective multicenter Phase 2 study assessing
safety and efficacy. J. Hepatol. 56(Suppl. 2):S554
104. Kingham TP, Karkar AM, D’Angelica MI, Allen PJ, DeMatteo RP, et al. 2012. Ablation of perivascular
hepatic malignant tumors with irreversible electroporation. J. Am. Coll. Surg. 215:379–87
105. Ball C, Thomson KR, Kavnoudias H. 2010. Irreversible electroporation: a new challenge in “out of
operating theater” anesthesia. Anesth. Analg. 110:1305–9
106. Akinwande O, Samman M. 2013. Irreversible electroporation: hype or hope? Cardiovasc. Interv. Radiol.
36:1719–20
107. Garcia PA, Pancotto T, Rossmeisl JH Jr, Henao-Guerrero N, Gustafson NR, et al. 2011. Non-thermal
irreversible electroporation (N-TIRE) and adjuvant fractionated radiotherapeutic multimodal therapy
for intracranial malignant glioma in a canine patient. Technol. Cancer Res. Treat. 10:73–83
108. Neal RE, Rossmeisl JH, Garcia PA, Lanz OI, Henao-Guerrero N, Davalos RV. 2011. Successful treatment of a large soft tissue sarcoma with irreversible electroporation. J. Clin. Oncol. 29:e372–77
109. Ellis TL, Garcia PA, Rossmeisl JH Jr, Henao-Guerrero N, Robertson J, Davalos RV. 2011. Nonthermal
irreversible electroporation for intracranial surgical applications: laboratory investigation. J. Neurosurg.
114:681–88
110. Heller LC, Heller R. 2010. Electroporation gene therapy preclinical and clinical trials for melanoma.
Curr. Gene Ther. 10:312–17
111. Čemažar M, Jarm T, Serša G. 2010. Cancer electrogene therapy with interleukin-12. Curr. Gene Ther.
10:300–11
112. Zhao Y, Moon E, Carpenito C, Paulos CM, Liu X, et al. 2010. Multiple injections of electroporated
autologous T cells expressing a chimeric antigen receptor mediate regression of human disseminated
tumor. Cancer Res. 70:9053–61
113. Sardesai NY, Weiner DB. 2011. Electroporation delivery of DNA vaccines: prospects for success. Curr.
Opin. Immunol. 23:421–29
114. Rols MP, Delteil C, Golzio M, Dumond P, Cros S, Teissié J. 1998. In vivo electrically mediated protein
and gene transfer in murine melanoma. Nat. Biotechnol. 16:168–71
115. Heller R, Jaroszeski M, Atkin A, Moradpour D, Gilbert R, et al. 1996. In vivo gene electroinjection and
expression in rat liver. FEBS Lett. 389:225–28
116. Aihara H, Miyazaki J. 1998. Gene transfer into muscle by electroporation in vivo. Nat. Biotechnol. 16:867–
70
117. Heller LC, Heller R. 2006. In vivo electroporation for gene therapy. Hum. Gene Ther. 17:890–97
118. Favard C, Dean DS, Rols MP. 2007. Electrotransfer as a non viral method of gene delivery. Curr. Gene
Ther. 7:67–77
www.annualreviews.org • Electroporation in Medicine
317
BE16CH12-Yarmush
ARI
19 June 2014
14:20
119. Čemažar M, Serša G. 2007. Electrotransfer of therapeutic molecules into tissues. Curr. Opin. Mol. Ther.
9:554–62
120. André F, Gehl J, Serša G, Préat V, Hojman P, et al. 2008. Efficiency of high- and low-voltage pulse
combinations for gene electrotransfer in muscle, liver, tumor, and skin. Hum. Gene Ther. 19:1261–71
121. Martel-Renoir D, Trochon-Joseph V, Galaup A, Bouquet C, Griscelli F, et al. 2003. Coelectrotransfer
to skeletal muscle of three plasmids coding for antiangiogenic factors and regulatory factors of the
tetracycline-inducible system: tightly regulated expression, inhibition of transplanted tumor growth,
and antimetastatic effect. Mol. Ther. 8:425–33
122. Jazowiecka-Rakus J, Jarosz M, Szala S. 2006. Combination of vasostatin gene therapy with cyclophosphamide inhibits growth of B16(F10) melanoma tumours. Acta Biochim. Pol. 53:199–202
123. Cichon T, Jamrozy L, Glogowska J, Missol-Kolka E, Szala S. 2002. Electrotransfer of gene encoding
endostatin into normal and neoplastic mouse tissues: inhibition of primary tumor growth and metastatic
spread. Cancer Gene Ther. 9:771–77
124. Tao J, Tu YT, Huang CZ, Feng AP, Wu Q, et al. 2005. Inhibiting the growth of malignant melanoma
by blocking the expression of vascular endothelial growth factor using an RNA interference approach.
Br. J. Dermatol. 153:715–24
125. Bošnjak M, Prosen L, Dolinšek T, Blagus T, Markelc B, et al. 2013. Biological properties of melanoma
and endothelial cells after plasmid AMEP gene electrotransfer depend on integrin quantity on cells.
J. Membr. Biol. 246:803–19
126. Dolinšek T, Markelc B, Serša G, Coer A, Štimac M, et al. 2013. Multiple delivery of siRNA against
endoglin into murine mammary adenocarcinoma prevents angiogenesis and delays tumor growth. PLoS
ONE 8:e58723
127. Spanggaard I, Snoj M, Cavalcanti A, Bouquet C, Serša G, et al. 2013. Gene electrotransfer of plasmid
antiangiogenic metargidin peptide (AMEP) in disseminated melanoma: safety and efficacy results of a
Phase I first-in-man study. Hum. Gene Ther. Clin. Dev. 24(3):99–107
128. Pavlin D, Čemažar M, Serša G, Tozon N. 2012. IL-12 based gene therapy in veterinary medicine.
J. Transl. Med. 10:234
129. Tamzali Y, Couderc B, Rols MP, Golzio M, Teissié J. 2007. Equine cutaneous tumors treatment by
electro-chemo-immuno-geno-therapy. In 11th Mediterr. Conf. Med. Biomed. Eng. Comput. 2007 Proc., ed.
T Jarm, P Kramar, A Zupanic, p. 630. Berlin: Springer
130. Reed SD, Fulmer A, Buckholz J, Zhang B, Cutrera J, et al. 2010. Bleomycin/interleukin-12 electrochemogenetherapy for treating naturally occurring spontaneous neoplasms in dogs. Cancer Gene Ther. 17:571–78
131. Nomura M, Nakata Y, Inoue T, Uzawa A, Itamura S, et al. 1996. In vivo induction of cytotoxic
T lymphocytes specific for a single epitope introduced into an unrelated molecule. J. Immunol. Methods
193:41–49
132. Van Drunen Littel-Van den Hurk S, Hannaman D. 2010. Electroporation for DNA immunization:
clinical application. Expert Rev. Vaccines 9:503–17
133. Livingston BD, Little SF, Luxembourg A, Ellefsen B, Hannaman D. 2010. Comparative performance
of a licensed anthrax vaccine versus electroporation based delivery of a PA encoding DNA vaccine in
rhesus macaques. Vaccine 28:1056–61
134. Best SR, Peng S, Juang CM, Hung CF, Hannaman D, et al. 2009. Administration of HPV DNA vaccine
via electroporation elicits the strongest CD8+ T cell immune responses compared to intramuscular
injection and intradermal gene gun delivery. Vaccine 27:5450–59
135. Rosati M, Valentin A, Jalah R, Patel V, Von Gegerfelt A, et al. 2008. Increased immune responses in
rhesus macaques by DNA vaccination combined with electroporation. Vaccine 26:5223–29
136. Hacein-Bey-Abina S, Von Kalle C, Schmidt M, Le Deist F, Wulffraat N, et al. 2003. A serious adverse
event after successful gene therapy for X-linked severe combined immunodeficiency. N. Engl. J. Med.
348:255–56
137. Ribeiro S, Mairhofer J, Madeira C, Diogo MM, Lobato da Silva C, et al. 2012. Plasmid DNA size does
affect nonviral gene delivery efficiency in stem cells. Cell. Reprogram. 14:130–37
138. Epsztejn-Litman S, Eiges R. 2010. Genetic manipulation of human embryonic stem cells. In Human
Embryonic Stem Cell Protocols, ed. K Turksen, pp. 387–411. New York: Humana
318
Yarmush et al.
BE16CH12-Yarmush
ARI
19 June 2014
14:20
139. Fernand F, Rubinsky L, Golberg A, Rubinsky B. 2012. Variable electric fields for high throughput
electroporation protocol design in curvilinear coordinates. Biotechnol. Bioeng. 109:2168–71
140. Shengnian W, Lee LJ. 2013. Micro-/nanofluidics based cell electroporation. Biomicrofluidics 7:011301
141. Huang Y, Rubinsky B. 2001. Microfabricated electroporation chip for single cell membrane permeabilization. Sens. Actuators A Phys. 89:242–49
142. Geng T, Zhan Y, Wang HY, Witting SR, Cornetta KG, Lu C. 2010. Flow-through electroporation
based on constant voltage for large-volume transfection of cells. J. Control. Release 144:91–100
143. Zorec B, Préat V, Miklavčič D, Pavšelj N. 2013. Active enhancement methods for intra- and transdermal
drug delivery: a review. Slovenian Med. J. 82:339–56
144. Denet A-R, Vanbever R, Préat V. 2004. Skin electroporation for transdermal and topical delivery. Adv.
Drug Deliv. Rev. 56:659–74
145. Prausnitz MR, Langer R. 2008. Transdermal drug delivery. Nat. Biotechnol. 26:1261–68
146. Pavšelj N, Préat V, Miklavčič D. 2007. A numerical model of skin electropermeabilization based on in
vivo experiments. Ann. Biomed. Eng. 35:2138–44
147. Pliquett U, Weaver JC. 2007. Feasibility of an electrode-reservoir device for transdermal drug delivery
by noninvasive skin electroporation. IEEE Trans. Biomed. Eng. 54:536–38
148. Pavšelj N, Miklavčič D. 2008. A numerical model of permeabilized skin with local transport regions.
IEEE Trans. Biomed. Eng. 55:1927–30
149. Vanbever R, LeBoulenge E, Préat V. 1996. Transdermal delivery of fentanyl by electroporation. I.
Influence of electrical factors. Pharm. Res. 4:559–65
150. Wong TW, Chen TY, Huang CC, Tsai JC, Hui SW. 2011. Painless skin electroporation as a novel way
for insulin delivery. Diabetes Technol. Ther. 13:929–35
151. Blagus T, Markelc B, Čemažar M, Kosjek T, Préat V, et al. 2013. In vivo real-time monitoring system of
electroporation mediated control of transdermal and topical drug delivery. J. Control. Release 172:862–71
152. Regnier V, DeMorre N, Jadoul A, Préat V. 1999. Mechanisms of a phosphorothioate oligonucleotide
delivery by skin electroporation. Int. J. Pharm. 184:147–56
153. Dujardin N, Van Deŕ Smissen P, Préat V. 2001. Topical gene transfer into rat skin using electroporation.
Pharm. Res. 1:61–66
154. Pavšelj N, Préat V. 2005. DNA electrotransfer into the skin using a combination of one high- and one
low-voltage pulse. J. Control. Release 106:407–15
155. Zhao YL, Murthy SN, Manjili MH, Guan LJ, Sen A, Hui SW. 2006. Induction of cytotoxic
T-lymphocytes by electroporation-enhanced needle-free skin immunization. Vaccine 24:1282–90
156. Kang TH, Monie A, Wu LSF, Pang X, Hung C-F, Wu TC. 2011. Enhancement of protein vaccine
potency by in vivo electroporation mediated intramuscular injection. Vaccine 29:1082–89
157. Edd JF, Davalos RV. 2007. Mathematical modeling of irreversible electroporation for treatment planning. Technol. Cancer Res. Treat. 6:275–86
158. Zupanic A, Kos B, Miklavčič D. 2012. Treatment planning of electroporation-based medical interventions: electrochemotherapy, gene electrotransfer and irreversible electroporation. Phys. Med. Biol.
57:5425–40
159. Garcia PA, Rossmeisl JH, Neal RE, Ellis TL, Olson JD, et al. 2010. Intracranial nonthermal irreversible
electroporation: in vivo analysis. J. Membr. Biol. 236:127–36
160. Appelbaum L, Ben-David E, Sosna J, Nissenbaum Y, Goldberg SN. 2012. US findings after irreversible
electroporation ablation: radiologic-pathologic correlation. Radiology 262:117–25
161. Wendler JJ, Pech M, Porsch M, Janitzky A, Fischbach F, et al. 2012. Urinary tract effects after multifocal nonthermal irreversible electroporation of the kidney: acute and chronic monitoring by magnetic
resonance imaging, intravenous urography and urinary cytology. Cardiovasc. Interv. Radiol. 35:921–26
162. Calmels L, Al-Sakere B, Ruaud J-P, Leroy-Willig A, Mir LM. 2012. In vivo MRI follow-up of murine
tumors treated by electrochemotherapy and other electroporation-based treatments. Technol. Cancer Res.
Treat. 11:561–70
163. Hjouj M, Last D, Guez D, Daniels D, Sharabi S, et al. 2012. MRI study on reversible and irreversible
electroporation induced blood brain barrier disruption. PLoS ONE 7:e42817
www.annualreviews.org • Electroporation in Medicine
319
BE16CH12-Yarmush
ARI
19 June 2014
14:20
164. Corovic S, Zupanic A, Kranjc S, Al Sakere B, Leroy-Willig A, et al. 2010. The influence of skeletal
muscle anisotropy on electroporation: in vivo study and numerical modeling. Med. Biol. Eng. Comput.
48:637–48
165. Neal R II, Cheung W, Kavnoudias H, Thomson K. 2012. Spectrum of imaging and characteristics for
liver tumors treated with irreversible electroporation. J. Biomed. Sci. Eng. 5:813–18
166. Mozzillo N, Caraco C, Mori S, Di Monta G, Botti G, et al. 2012. Use of neoadjuvant electrochemotherapy
to treat a large metastatic lesion of the cheek in a patient with melanoma. J. Transl. Med. 10:131
167. Cukjati D, Batiuskaite D, André F, Miklavčič D, Mir LM. 2007. Real time electroporation control for
accurate and safe in vivo non-viral gene therapy. Bioelectrochemistry 70:501–7
168. Golberg A, Laufer S, Rabinowitch HD, Rubinsky B. 2011. In vivo non-thermal irreversible electroporation impact on rat liver galvanic apparent internal resistance. Phys. Med. Biol. 56:951–63
169. Granot Y, Ivorra A, Maor E, Rubinsky B. 2009. In vivo imaging of irreversible electroporation by means
of electrical impedance tomography. Phys. Med. Biol. 54:4927–43
170. Ivorra A, Al-Sakere B, Rubinsky B, Mir LM. 2009. In vivo electrical conductivity measurements during
and after tumor electroporation: Conductivity changes reflect the treatment outcome. Phys. Med. Biol.
54:5949–63
171. Golberg A, Rabinowitch HD, Rubinsky B. 2009. Galvanic apparent internal impedance: an intrinsic
tissue property. Biochem. Biophys. Res. Commun. 389:168–71
172. Granot Y, Rubinsky B. 2007. Methods of optimization of electrical impedance tomography for imaging
tissue electroporation. Physiol. Meas. 28:1135–47
173. Kranjc M, Bajd F, Serša I, Miklavčič D. 2011. Magnetic resonance electrical impedance tomography for
monitoring electric field distribution during tissue electroporation. IEEE Trans. Med. Imaging 30:1771–
78
174. Kranjc M, Bajd F, Serša I, Woo EJ, Miklavčič D. 2012. Ex vivo and in silico feasibility study of monitoring
electric field distribution in tissue during electroporation based treatments. PLoS ONE 7:e45737
175. Pavliha D, Kos B, Marcan M, Zupanic A, Serša G, Miklavčič D. 2013. Planning of electroporation-based
treatments using web-based treatment-planning software. J. Membr. Biol. 246:833–42
176. Arena CB, Sano MB, Rossmeisl JH, Caldwell JL, Garcia PA, et al. 2011. High-frequency irreversible
electroporation (H-FIRE) for non-thermal ablation without muscle contraction. Biomed. Eng. OnLine
10:102
177. Golberg A, Rubinsky B. 2012. Towards electroporation based treatment planning considering electric
field induced muscle contractions. Technol. Cancer Res. Treat. 11:189–201
178. Sano M, Neal R, Garcia P, Gerber D, Robertson J, Davalos R. 2010. Towards the creation of decellularized organ constructs using irreversible electroporation and active mechanical perfusion. Biomed. Eng.
OnLine 9:83
179. Markelc B, Tevž G, Čemažar M, Kranjc S, Lavrenčak J, et al. 2012. Muscle gene electrotransfer is
increased by the antioxidant tempol in mice. Gene Ther. 19:312–20
180. Čemažar M, Golzio M, Serša G, Escoffre JM, Coer A, et al. 2012. Hyaluronidase and collagenase increase
the transfection efficiency of gene electrotransfer in various murine tumors. Hum. Gene Ther. 23:128–37
181. Kotnik T. 2013. Lightning-triggered electroporation and electrofusion as possible contributors to natural
horizontal gene transfer. Phys. Life Rev. 10:351–70
182. Dev SB, Rabussay DP, Widera G, Hofmann GA. 2000. Medical applications of electroporation. IEEE
Trans. Plasma Sci. 28:206–23
183. Bower M, Sherwood L, Li Y, Martin R. 2011. Irreversible electroporation of the pancreas: definitive
local therapy without systemic effects. J. Surg. Oncol. 104:22–28
184. Zorec B, Miklavčič D, Jelenc J, Pavšelj N. 2013. Enhanced transport of Patent Blue dye through the
stratum corneum using Green Skin Pore electroporator. J. Laser Health Acad. 1:S30
320
Yarmush et al.